Next Article in Journal
Progress in Studies of Climatic Suitability of Crop Quality and Resistance Mechanisms in the Context of Climate Warming
Next Article in Special Issue
Population Structure and Genetic Diversity of Colletotrichum gloeosporioides on Citrus in China
Previous Article in Journal
Variations of Rice Yield and Quality in Response to Different Establishment Methods at Farmers’ Field
Previous Article in Special Issue
Potential of Omics to Control Diseases and Pests in the Coconut Tree
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

ACT-Toxin, the Key Effector for the Virulence of Alternaria alternata Tangerine Pathotype to Specific Citrus Species

1
Collaborative Innovation Center for Efficient and Green Production of Agriculture in Mountainous Areas of Zhejiang Province, College of Horticulture Science, Zhejiang A&F University, Hangzhou 311300, China
2
School of Grassland Science, Beijing Forestry University, Beijing 100083, China
3
Key Lab of Molecular Biology of Crop Pathogens and Insects, Ministry of Agriculture, Institute of Biotechnology, Zhejiang University, Hangzhou 310027, China
4
China-USA Citrus Huanglongbing Joint Laboratory (A Joint Laboratory of the University of Florida’s Institute of Food and Agricultural Sciences and Gannan Normal University), National Navel Orange Engineering Research Center, Gannan Normal University, Ganzhou 341000, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Agronomy 2022, 12(12), 3181; https://doi.org/10.3390/agronomy12123181
Submission received: 11 November 2022 / Revised: 6 December 2022 / Accepted: 8 December 2022 / Published: 15 December 2022
(This article belongs to the Special Issue Research Progress on Pathogenicity of Fungus in Crop)

Abstract

:
Alternaria brown spot disease is caused by the Alternaria alternata tangerine pathotype, which relies on ACT-toxin for infection. At present, all identified ACT-toxin biosynthesis-related genes are multi-copy genes. In this study, we summarized the advances in important host-specific toxins (HSTs), and listed key genes required for the pathogenicity of the A. alternata tangerine pathotype. Toxin virulence test results revealed that different citrus species displayed distinctly different tolerances to ACT-toxin. The extraction method of ACT-toxin crude extract was described in schematic form to make the method easier to understand. In addition, target gene disruption of two copies of ACTT5 (∆∆ACTT5) displayed significantly reduced virulence, indicating that ACTT5 is essential for the pathogenicity of the A. alternata tangerine pathotype.

1. Introduction

Host-specific toxins (HSTs) are essential for the pathogenicity of the corresponding phytopathogens [1]. Currently, numerous pathogens have been documented to produce HSTs, such as ACT toxin (A. alternata) [2], ABR toxin (Alternaria brassicae) [3], PC toxin (Periconia circinata) [4], HMT toxin (Helminthosporium maydis) [5], etc. HSTs range from low-molecular-weight metabolites to proteins, which are essential for the virulence of corresponding pathogens. However, the action mode of most HSTs remains unknown. Most HST-producing pathogens are necrotrophic or hemibiotrophic fungi. Alternaria species, which cause pathogenic disease on various economically important crops, are well-known producers of HSTs [6,7,8,9,10].
At present, more than 70 toxins with different chemical structures are known to be biosynthesized by Alternaria species [11], including Alternaria alternata, Alternaria arborescens, Alternaria brassicae, Alternaria brassicola, Alternaria infectoria, and Alternaria radicina [1,12,13]. The corresponding HSTs are toxic and could induce cell death, specifically at 10−9 to 10−8 M when applied to susceptible species [14]. The targets of these toxins include the plasma membrane, mitochondria, chloroplast, Golgi bodies, nucleus, etc. Alternaria species cause disease in about 400 plant species, of which A. alternata infects nearly 100 plant species [1]. Currently, 13 HSTs have been identified in Alternaria species, and most of them are produced by A. alternata [1]. Different pathotypes of A. alternata are determined by the toxins they produce. Chemical structures of at least six A. alternata HSTs have been determined [15]. For example, the A. alternata tangerine pathotype produces ACT-toxin [2], the Japanese pear pathotype produces AK-toxin [16], the strawberry pathotype produces AF-toxin [17], the tomato pathotype produces AAL-toxin [18], rough lemon produces ACR-toxin [19], etc. HSTs of tangerine, strawberry, and Japanese pear pathotypes were found to be structurally analogous metabolites that are esters of 9,10-epoxy-8-hydroxy-9-methyl-decatrienoic acid (EDA). The primary site of ACT-toxin, which contains three components, EDA, polyketide, and valine, is the plasma membrane [20]. Most of the genes required for HSTs biosynthesis in A. alternata pathotypes are multi-copy genes, and disruption of any of these genes results in the loss of virulence [21,22,23,24,25].
Alternaria brown spot (ABS), which is a severe fungal disease resulting in defoliation and fruit drop of citrus, is caused by the fungus A. alternata tangerine pathotype [26]. This fungus is strictly pathogenic as a result of its capability to produce the HST, ACT-toxin [27]. All the mutants that fail to produce ACT-toxin are nonpathogenic [24,28]. To date, ACTT2 [25], ACTTS2 [22], ACTT3 [27], ACTTS3 [27], ACTT5 [24], ACTT6 [24], and ACTTR [28] encoding genes have been revealed to be required for ACT-toxin biosynthesis, and are located in the ACT-toxin gene cluster of the conditionally dispensable chromosome (CDC) [29]. Genetic inactivation of any ACTT gene will block ACT-toxin biosynthesis and lead to the complete loss of virulence [22,27]. Simultaneous treatment with HSTs and non-pathogenic A. alternata strains resulted in successful infection [30,31].
In addition to ACT-toxin-biosynthesis-related genes, other genes involved in ROS detoxification or cell wall degradation are also required for the virulence of the A. alternata tangerine pathotype. For example, nicotinamide adenine dinucleotide phosphate oxidase genes are required for the accumulation of cellular hydrogen peroxide (H2O2), reactive oxygen species (ROS) detoxification, and pathogenicity [32,33]. In addition, other genes (AP1, Hog1, Skn7, Tsa1, etc.) involved in ROS detoxification are also required for the virulence of A. alternata [34,35,36,37]. On the other hand, the capability to break through the citrus cuticle layer via cell-wall-degrading enzymes is also required for the full virulence of A. alternata [38,39].
In this study, to demonstrate whether the susceptibility of citrus to A. alternata is determined by the tolerance level to ACT-toxin, citrus species, which are resistant, tolerant, and susceptible to ABS, were analyzed. In addition, we performed pathogenicity analysis of mutants with one (ACTT5) ACT-toxin-biosynthesis-related gene disrupted on leaves of different citrus species, as all these genes are essential for ACT-toxin biosynthesis and are not involved in vegetative growth and conidiation according to previous study [22,24,25,27,28].

2. Materials and Methods

2.1. Strains and Plants

The wild-type strain of A. alternata Z7 (CGMCC3.18907) was isolated from infected tangerine ‘Ougan’ [26,29,40]. All strains were grown on potato dextrose agar medium (PDA), potato dextrose broth medium (PDB), regeneration medium (RMM), or V8 medium at 26 °C. Citrus leaves were collected from Dancy (Citrus reticulata), Clementina (Citrus clementina), Minneola, sweet orange (C. sinensis), Citrus medica, Carrizo citrange, and Sugar belle. All the citrus plants provided by the National Citrus Engineering Research Center or Fred G. Gmitter, Jr were grown in a greenhouse.

2.2. Medium or Solution

PDA medium: 20 g glucose, 200 g potato, 20 g agar, add ddH2O to 1 L.
PDB medium: 20 g glucose, 200 g potato, add ddH2O to 1 L.
Trace elements solution: 5 g citric acid, 5 g ZnSO4, 1 g (NH4)2Fe(SO4)2·6H2O, add ddH2O to 100 mL.
V8 medium: 200 mL V8 broth, 3 g CaCO3, 20 g agar, add ddH2O to 1 L.
Richards’ medium: 25 g glucose, 10 g KNO3, 5 g KH2PO4, 2.5 g MgSO4, 0.02 g FeCl3, and 0.005 g ZnSO4, add ddH2O to 1 L.

2.3. Conidiation

Z7 and corresponding mutants were incubated on V8 medium at 26 °C for 8 days. Then, the conidia were harvested from colonies grown on V8 medium by immersion, scraped with sterile water, and passed through three layers of sterile cheesecloth.

2.4. ACTT5 Knockout

Previous studies have revealed that ACTT5 contains at least three copies [24]. Two copies of ACTT5 (ACTT5-1 and ACTT5-2) were disrupted in Z7 strain using a homologous recombination method in this study. First, for ACTT5-1 disruption in Z7 to obtain ∆ACTT5, a fragment containing HPH (phosphotransferase) encoding gene, 5′ACTT5-1 fragment (1250 bp length, 2021 bp before the start codon of ACTT5-1), and 3′ACTT5-1 fragment (1032 bp length, 1787 bp after the stop codon of ACTT5-1) was constructed. The fragment was transformed into protoplasts prepared from Z7 strain using CaCl2 and polyethylene glycol, as described by Chung et al. [41]. Fungal transformants were recovered in a regeneration medium amended with 200 µg/mL hygromycin (Roche Applied Science, Indianapolis, IN, USA). Then, for ACTT5-2 disruption in ∆ACTT5 to obtain ∆∆ACTT5, a fragment containing Neo encoding gene, 5′ACTT5-2 fragment (914 bp length, 210 bp before the start codon of ACTT5-2), and 3′ACTT5-2 fragment (1096 bp length, 381 bp after the stop codon of ACTT5-2) was constructed. The fragment was transformed into protoplasts prepared from Z7 strain using CaCl2 and polyethylene glycol. Fungal transformants were recovered in a regeneration medium amended with 100 µg/mL G418. The corresponding gene knockout mutants were verified based on the phenotype according to a previous study [24].

2.5. Determination of Pathogenicity

Pathogenicity assay of corresponding strains was performed on detached leaves of Dancy, Clementina, Minneola, and Sugar belle with conidial suspension (1 × 104 conidia/mL). Conidial suspension was sprayed on tested leaves and the inoculated leaves were incubated in a mist chamber for 5 to 8 days for lesion development. Each fungal strain was tested on at least 10 leaves, and experiments were repeated twice.

2.6. ACT-Toxin Extraction and Virulence Assay

Z7 strain was grown in 200 mL of modified Richards’ medium [2,42,43] at 26 °C for 25–35 days, and the culture filtrate was collected through three layers of gauze. The culture filtrate was adjusted to pH 5.5 with 10% NaH2PO4 and was stirred with 1 L of Amberlite XAD-2 resin (Aldrich Chemical Co., Inc., Milwaukee, WI, USA) for 2 h to absorb toxins. The XAD-2 was packed in a funnel and eluted with methanol. The eluate was concentrated using a rotary evaporator. The virulence assay of ACT-toxin crude extract was performed by using resistant (C. clementina, C. medica, C. sinensis, and Carrizo citrange) and susceptible (C. reticulata) leaves. The petioles of tested citrus leaves were inserted in PCR tubes containing ACT-toxin crude extract at 26 °C for 1–2 days. The phenotype appearing around petioles was recorded after incubation.

2.7. Phenotypic Analysis

Stress tolerance was assayed by placing mycelia plugs (5 mm × 5 mm) onto PDA or MM medium amended with oxidants, salts, cell wall-interfering agents, DNA-damaging agents, or other indicated chemicals at 26 °C for 3–5 days. The percentage of growth reduction was determined by comparing a cumulative percentage of the growth of Z7 and mutants grown on same medium. All tests were repeated at least twice with three replicates of each treatment. Tested compounds included sorbitol (1 mM), NaCl (1 M), H2O2 (10 mM), sodium dodecyl sulfate (SDS, 0.01%), and Congo red (CR, 0.2 mg/mL).

2.8. Statistical Analysis

The diameter of growth is presented as the mean ± standard deviation (SD) (n = 6). Significance was determined using Student’s t-test (p < 0.05) for differences between the wild-type and ∆∆ACTT5.

3. Results

3.1. ACT-Toxin Biosynthesis and ROS Detoxification-Related Genes Are Involved in the Pathogenicity of A. alternata

HSTs produced by fungal plant pathogens are key factors responsible for pathogenicity during certain plant-pathogen interactions. Table 1 shows some important HSTs produced by phytopathogenic fungi. Nine different pathotypes of A. alternata known to produce HSTs are listed in Table 2. ACT-toxin, which is produced by the A. alternata tangerine pathotype, is essential for the pathogenesis of this pathogen. ACT-toxin biosynthesis is regulated by multiple genes in the ACT-toxin gene cluster, and the knockout of any one of these genes results in the complete loss of pathogenicity. Therefore, we listed relevant information (including accession number, copy number, and corresponding references) of these genes in Table 3. In addition to ACT-toxin, genes involved in ROS detoxification are also required for the virulence of A. alternata. Table 4 shows the studied genes required for ROS detoxification and pathogenicity in A. alternata.

3.2. Different Citrus Species Display Distinct Sensitivity to ACT-Toxin

Until now, no schematic diagram of the process for obtaining ACT-toxin crude extract has been available, although ACT-toxin is crucial for the analysis of the pathogenicity mechanism in A. alternata. Therefore, we made a schematic diagram of ACT-toxin extraction. Briefly, the extraction protocol comprised five steps: toxin-producing cultivation, culture filtrate obtaining, toxin adsorption using XAD-2, toxin dissolution, and toxin concentration (Figure 1A). Subsequently, we obtained ACT-toxin crude extract from Z7 strain using this protocol. ACT-toxin virulence assay was performed on detached leaves of C. reticulata, C. clementina, C. medica, C. sinensis, and Carrizo citrange. Necrotic lesions quickly developed on the leaves of C. reticulata. On the contrary, C. clementina, C. medica, C. sinensis, and Carrizo citrange, which are citrus species resistant to ABS, displayed highly tolerance to ACT-toxin (Figure 1B). These results revealed that the ABS resistance capability of citrus was derived from the tolerance capability to ACT-toxin.

3.3. ACTT5 Is Required for the Pathogenicity of A. alternata

To verify that the sensitivity of citrus to ACT-toxin is the major factor in determining its resistance or susceptibility capability to ABS, we knocked out two copies of ACTT5 using the developed multi-copy gene disruption strategy (Figure 2A, Table S1). ∆ACTT5 means mutant with one copy of ACTT5 knocked out. ∆∆ACTT5 means mutant with two copies of ACTT5 knocked out. Both ∆ACTT5 and ∆∆ACTT5 displayed wild-type radial growth on PDA medium (Figure 2B), indicating ACTT5 was not involved in the vegetative growth of A. alternata. Pathogenicity analysis revealed that the wild-type Z7 strain could induce significantly enlarged necrotic lesions on leaves of ABS-susceptible citrus species (Dancy and Minneola), and tiny necrotic lesions on ABS-tolerant species (Sugar belle). No obvious necrotic lesions were observed on ABS-resistant accession (C. clementina) (Figure 2C). ∆∆ACTT5 did not produce any visible lesions on ABS-resistant and -tolerant species (C. clementina and Sugar belle), and induced almost undetectable tiny lesions on susceptible citrus species (Dancy and Minneola) (Figure 2C).

3.4. ACTT5 Is Not Involved in Conidiation, Vegetative Growth, and Multi-Stress Resistance

The ∆∆ACTT5 produced regular mycelium and ovoid conidia with dark pigmentation, similar to those produced by the wild-type, indicating ACTT5 is not involved in cell development (Figure 3A,B). ∆∆ACTT5 displayed wild-type sensitivity to sorbitol and NaCl, indicating that ACTT5 is not required for the resistance to osmotic stress. In addition, ∆∆ACTT5 also displayed wild-type sensitivity to H2O2, SDS, and Congo red (CR), indicating ACTT5 is not involved in ROS detoxification and cell wall integrity (Figure 3C).

4. Discussion

HSTs are a group of structurally complex and chemically diverse metabolites produced by specific phytopathogenic fungi and function as essential determinants of virulence. HSTs-producing species mainly include Alternaria, Helminthosporium, Colletotrichum, Fusarium, Periconia, Phyllosticta, Corynespora, etc. Alternaria species cause pathogenic disease on numerous economic crops. A. alternata consists of at least seven pathotypes, each of which can produce a specific HST toxic in the corresponding host plant [15]. Citrus is the world’s largest type of fruit. The A. alternata tangerine pathotype, which produces ACT-toxin, causes ABS in tangerines (C. reticulata), grapefruits, (C. paradisi Macfad), hybrids of tangerine and sweet orange, and hybrids of tangerine and grapefruit [40,68]. Previous studies have revealed that ACT-toxin can cause necrotic lesions on citrus leaves at a concentration of 2 × 10−8 M [2,69]. Therefore, the study of the interaction mechanism of ACT-toxin and citrus is of great significance for the breeding of ABS-resistant citrus varieties. However, research on the action mechanism of ACT-toxin is still at the stage of microscopic observation. In this study, we summarized 17 HSTs produced by different pathogens, provided the details of 7 HSTs produced by A. alternata, and listed studied genes essential for the biosynthesis of ACT-toxin.
Previous studies have revealed that ACT-toxin biosynthesis and ROS detoxification capability is essential for the pathogenicity of A. alternata. ROS is generated from host cells during the infection of A. alternata [70]. A. alternata must be able to perform ROS detoxification for successful colonization. Therefore, many key genes essential for ROS detoxification were also revealed to be needed for the pathogenicity, such as Ap1 [62], Hog1 [35], Skn7 [36], etc. To date, the identified genes responsible for ACT-toxin biosynthesis include ACTT1, ACTT2, ACTTS2, ACTT3, ACTTS3, ACTT5, ACTT6, ACTTR, etc. [18,22,24,25,27,32]. These genes, which are mainly clustered and located in the conditionally dispensable chromosome, are multi-copy genes [29]. Among them, ACTT1 and ACTT2 are considered as homologous genes of AKT1 and AKT2 of the A. alternata Japanese pear pathotype [24]. ACTTS2 (enoyl reductase) and ACTTS3 (peptide synthetase) are also involved in ACT-toxin biosynthesis [22,27]. In addition, the corresponding homologous genes of ACTT5 (acetyl-CoA synthetase) and ACTT6 (enoyl-CoA hydratase) also exist in the Japanese pear pathotype. ACTTR encoding Zn2Cys6 transcription factor has been proven to be involved in ACT-toxin biosynthesis and pathogenicity in the A. alternata tangerine pathotype [28]. In this study, we knocked out two copies of ACTT5. We did not conduct an RNA silencing experiment for ACTT5 because the phenotype of ΔΔACTT5 met our experimental requirements. The virulence of ∆∆ACTT5 significantly decreased. Moreover, ∆∆ACTT5 displayed a wild-type phenotype of vegetative growth, conidiation, and multi-stress resistance. All these results revealed that ACT-toxin is a key factor for the virulence of the A. alternata tangerine pathotype.
Generally, the plasma membrane, mitochondria, chloroplast, and some important enzymes are the inhibitory sites for the action of HSTs of A. alternata [1,15]. Based on the chemical structures of HSTs produced by A. alternata pathotypes, HSTs can be classified into seven classes. The structure of ACT-toxin is related to AF- and AK-toxins, each of which share a common EDA moiety [2,17,44]. The structure of the other HSTs is as follows: AAL-toxin (sphinganine analogue), ACR-toxin (pyranones), AM-toxin (cyclic peptide), AS-I toxin (tetrapeptide), and maculosin (diketopiperazine) [71]. The target of ACT-, AK-, and AF-toxins is the plasma membrane [69]. ACR-toxin first targets the mitochondria and then other cell organelles [69]. Mitochondria and endoplasmic reticulum are the primary targets for the action of AAL-toxin. Chloroplasts are the primary target of AM-toxin and maculosin. The A. alternata tangerine pathotype produces ACT-toxins I and II, with ACT-toxin I being more toxic to citrus cells [2]. The ABS susceptibility of citrus species has been studied by numerous researchers [72,73,74,75]. In this study, we provided a schematic diagram for obtaining ACT-toxin crude extract. ACT-toxin inoculation results revealed that ABS-resistant citrus accession (C. clementina) was tolerant to ACT-toxin, and the ABS-susceptible citrus accession (C. reticulata) was highly sensitive to ACT-toxin. The disruption of the essential gene (ACTT5) involved in ACT-toxin biosynthesis resulted in the loss of virulence for A. alternata. All these results revealed that the ABS resistance capability of citrus is dependent on the ACT-toxin tolerance capability.

5. Conclusions

In this study, we reviewed HSTs produced by phytopathogenic fungi and summarized the HSTs of different A. alternata pathotypes. Our results further proved that ACT-toxin tolerance capability is an important basis on which to analyze whether the tested citrus species are resistant to ABS.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/agronomy12123181/s1, Table S1: Primers used in this study.

Author Contributions

Conceptualization, H.M.; writing, H.M. and Y.G.; Supervision, H.M.; experiment, S.H., Z.J., H.L., S.Z. and J.S.; Plants maintenance, C.J., X.S., M.W. and S.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (32202427), the Key Project for New Agricultural Cultivar Breeding in Zhejiang Province, China (2021C02066-1), as well as the Project Supported by Natural Science Foundation of Jiangxi, China (20202BABL215013).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Meena, M.; Gupta, S.K.; Swapnil, P.; Zehra, A.; Dubey, M.K.; Upadhyay, R.S. Alternaria toxins: Potential virulence factors and genes related to pathogenesis. J. Front. Microbiol. 2017, 8, 1451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Kohmoto, K.; Itoh, Y.; Shimomura, N.; Shimomura, Y.; Kondoh, Y.; Otani, H.; Kodama, M.; Nishimura, S.; Nakatsuka, S. Isolation and biological activities of two host-bpecific toxins from the tangerine pathotype of Alternaria alternata. Phytopathology 1993, 83, 495–502. [Google Scholar] [CrossRef]
  3. Parada, R.; Sakuno, E.; Mori, N.; Oka, K.; Egusa, M.; Kodama, M.; Otani, H. Alternaria brassicae produces a host-specific protein toxin from germinating spores on host leaves. Phytopathology 2008, 98, 458–463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kinoshita, T.; Kono, Y.; Takeuchi, S.; Daly, J.M. Structure of HV-toxin M, a host-specific toxin-related compound produced by Helminthosporium victoriae. Agric. Biol. Chem. 1989, 53, 1283–1290. [Google Scholar] [CrossRef]
  5. Holden, M.J.; Sze, H. Helminthosporium maydis T toxin increased membrane permeability to Ca2+ in susceptible corn mitochondria. Plant Physiol. 1984, 75, 235–237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Nishimura, S.; Kohmoto, K. Host-specific toxins and chemical structures from Alternaria species. Annu. Rev. Phytopathol. 1983, 21, 87–116. [Google Scholar] [CrossRef] [PubMed]
  7. Scheffer, R.P.; Livingston, R.S. Host-selective toxins and their role in plant diseases. Science 1984, 223, 17–21. [Google Scholar] [CrossRef]
  8. Walton, J.D. Host-selective toxins: Agents of compatibility. Plant Cell 1996, 8, 1723. [Google Scholar]
  9. Wolpert, T.J.; Dunkle, L.D.; Ciuffetti, L.M. Host-selective toxins and avirulence determinants: What’s in a name? Annu. Rev. Phytopathol. 2002, 40, 251–285. [Google Scholar] [CrossRef] [Green Version]
  10. Yoder, O. Toxins in pathogenesis. Annu. Rev. Phytopathol. 1980, 18, 103–129. [Google Scholar] [CrossRef]
  11. Lou, J.; Fu, L.; Peng, Y.; Zhou, L. Metabolites from Alternaria fungi and their bioactivities. Molecules 2013, 18, 5891–5935. [Google Scholar] [CrossRef]
  12. Thomma, B.P. Alternaria spp.: From general saprophyte to specific parasite. Mol. Plant Pathol. 2003, 4, 225–236. [Google Scholar] [CrossRef] [PubMed]
  13. Lee, H.B.; Patriarca, A.; Magan, N. Alternaria in food, Ecophysiology, mycotoxin production and toxicology. Mycobiology 2015, 43, 93–106. [Google Scholar] [CrossRef] [Green Version]
  14. Tanahashi, M.; Nakano, T.; Akamatsu, H.; Kodama, M.; Otani, H.; Osaki-Oka, K. Alternaria alternata apple pathotype (A. mali) causes black spot of European pear. Eur. J. Plant Pathol. 2016, 145, 787–795. [Google Scholar] [CrossRef]
  15. Tsuge, T.; Harimoto, Y.; Akimitsu, K.; Ohtani, K.; Kodama, M.; Akagi, Y.; Egusa, M.; Yamamoto, M.; Otani, H. Host-selective toxins produced by the plant pathogenic fungus Alternaria alternata. FEMS Microbiol. Rev. 2013, 37, 44–66. [Google Scholar] [CrossRef] [PubMed]
  16. Tanaka, A.; Shiotani, H.; Yamamoto, M.; Tsuge, T. Insertional mutagenesis and cloning of the genes required for biosynthesis of the host-specific AK-toxin in the Japanese pear pathotype of Alternaria alternata. Mol. Plant-Microbe Interact. 1999, 12, 691–702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Nakatsuka, S.; Ueda, K.; Goto, T.; Yamamoto, M.; Nishimura, S.; Kohmoto, K. Structure of AF-toxin II, one of the host-specific toxins produced by Alternaria alternata strawberry pathotype. Tetrahedron Lett. 1986, 27, 2753–2756. [Google Scholar] [CrossRef]
  18. Gilchrist, D.; Grogan, R. Production and nature of a host-specific toxin from Alternaria alternata f. sp. lycopersici. Phytopathology 1976, 66, 165–171. [Google Scholar] [CrossRef]
  19. Izumi, Y.; Ohtani, K.; Miyamoto, Y.; Masunaka, A.; Fukumoto, T.; Gomi, K.; Tada, Y.; Ichimura, K.; Peever, T.L.; Akimitsue, K. A polyketide synthase gene, ACRTS2, is responsible for biosynthesis of host-selective ACR-toxin in the rough lemon pathotype of Alternaria alternata. Mol. Plant-Microbe Interact. 2012, 25, 1419–1429. [Google Scholar] [CrossRef] [Green Version]
  20. Miyamoto, Y.; Masunaka, A.; Tsuge, T.; Yamamoto, M.; Ohtani, K.; Fukumoto, T.; Gomi, K.; Peever, T.L.; Akimitsu, K. Functional analysis of a multicopy host-selective ACT-toxin biosynthesis gene in the tangerine pathotype of Alternaria alternata using RNA silencing. Mol. Plant-Microbe Interact. 2008, 21, 1591–1599. [Google Scholar] [CrossRef] [Green Version]
  21. Akimitsu, K.; Tsuge, T.; Kodama, M.; Yamamoto, M.; Otani, H. Alternaria host-selective toxins: Determinant factors of plant disease. J. Gen. Plant Pathol. 2014, 80, 109–122. [Google Scholar] [CrossRef]
  22. Ajiro, N.; Miyamoto, Y.; Masunaka, A.; Tsuge, T.; Yamamoto, M.; Ohtani, K.; Fukumoto, T.; Gomi, K.; Peever, T.L.; Izumi, Y.; et al. Role of the host-selective ACT-toxin synthesis gene ACTTS2 encoding an enoyl-reductase in pathogenicity of the tangerine pathotype of Alternaria alternata. Phytopathology 2010, 100, 120–126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Izumi, Y.; Kamei, E.; Miyamoto, Y.; Ohtani, K.; Masunaka, A.; Fukumoto, T.; Gomi, K.; Tada, Y.; Ichimura, K.; Peever, T.L.; et al. Role of the pathotype-specific ACRTS1 gene encoding a hydroxylase involved in the biosynthesis of host-selective ACR-toxin in the rough lemon pathotype of Alternaria alternata. Phytopathology 2012, 102, 741–748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Miyamoto, Y.; Ishii, Y.; Honda, A.; Masunaka, A.; Tsuge, T.; Yamamoto, M.; Ohtani, K.; Fukumoto, T.; Gomi, K.; Peever, T.L.; et al. Function of genes encoding acyl-CoA synthetase and enoyl-CoA hydratase for host-selective ACT-toxin biosynthesis in the tangerine pathotype of Alternaria alternata. Phytopathology 2009, 99, 369–377. [Google Scholar] [CrossRef] [Green Version]
  25. Masunaka, A. Studies on genes for biosynthesis of host-selective ACT-toxin in the tangerine pathotype of Alternaria alternata. J. Gen. Plant Pathol. 2007, 73, 424–425. [Google Scholar] [CrossRef]
  26. Wang, M.; Sun, X.; Yu, D.; Xu, J.; Chung, K.; Li, H. Genomic and transcriptomic analyses of the tangerine pathotype of Alternaria alternata in response to oxidative stress. Sci. Rep. 2016, 6, 32437. [Google Scholar] [CrossRef] [Green Version]
  27. Miyamoto, Y.; Masunaka, A.; Tsuge, T.; Yamamoto, M.; Ohtani, K.; Fukumoto, T.; Gomi, K.; Peever, T.L.; Tada, Y.; Ichimura, K.; et al. ACTTS3 encoding a polyketide synthase is essential for the biosynthesis of ACT-toxin and pathogenicity in the tangerine pathotype of Alternaria alternata. Mol. Plant-Microbe Interact. 2010, 23, 406–414. [Google Scholar] [CrossRef] [Green Version]
  28. Li, L.; Ma, H.; Zheng, F.; Chen, Y.; Wang, M.; Jiao, C.; Li, H.; Gai, Y. The transcription regulator ACTR controls ACT-toxin biosynthesis and pathogenicity in the tangerine pathotype of Alternaria alternata. Microbiol. Res. 2021, 248, 126747. [Google Scholar] [CrossRef]
  29. Gai, Y.; Ma, H.; Chen, Y.; Li, L.; Cao, Y.; Wang, M.; Sun, X.; Jiao, C.; Riely, B.K.; Li, H.; et al. Chromosome-scale genome sequence of Alternaria alternata causing Alternaria brown spot of citrus. Mol. Plant-Microbe Interact. 2021, 34, 726–732. [Google Scholar] [CrossRef]
  30. Akimitsu, K.; Kohmoto, K.; Otani, H.; Nishimura, S. Host-specific effects of toxin from the rough lemon pathotype of Alternaria alternata on mitochondria. Plant Physiol. 1989, 89, 925–931. [Google Scholar] [CrossRef] [Green Version]
  31. Yamagishi, D.; Akamatsu, H.; Otani, H.; Kodama, M. Pathological evaluation of host-specific AAL-toxins and fumonisin mycotoxins produced by Alternaria and Fusarium species. J. Gen. Plant Pathol. 2006, 72, 323–327. [Google Scholar] [CrossRef]
  32. Yang, S.L.; Chung, K.R. Similar and distinct roles of NADPH oxidase components in the tangerine pathotype of Alternaria alternata. Mol. Plant Pathol. 2013, 14, 543–556. [Google Scholar] [CrossRef] [PubMed]
  33. Chung, K.R. Reactive oxygen species in the citrus fungal pathogen Alternaria alternata: The roles of NADPH-dependent oxidase. Physiol. Mol. Plant Pathol. 2014, 88, 10–17. [Google Scholar] [CrossRef]
  34. Lin, C.H.; Yang, S.L.; Wang, N.Y.; Chung, K.R. The FUS3 MAPK signaling pathway of the citrus pathogen Alternaria alternata functions independently or cooperatively with the fungal redox-responsive AP1 regulator for diverse developmental, physiological and pathogenic processes. Fungal Genet. Biol. 2010, 47, 381–391. [Google Scholar] [CrossRef]
  35. Lin, C.H.; Chung, K.R. Specialized and shared functions of the histidine kinase-and HOG1 MAP kinase-mediated signaling pathways in Alternaria alternata, a filamentous fungal pathogen of citrus. Fungal Genet. Biol. 2010, 47, 818–827. [Google Scholar] [CrossRef]
  36. Chen, L.H.; Lin, C.H.; Chung, K.R. Roles for SKN7 response regulator in stress resistance, conidiation and virulence in the citrus pathogen Alternaria alternata. Fungal Genet. Biol. 2012, 49, 802–813. [Google Scholar] [CrossRef]
  37. Ma, H.; Wang, M.; Gai, Y.; Fu, H.; Zhang, B.; Ruan, R.; Chung, K.-R.; Li, H. Thioredoxin and glutaredoxin systems required for oxidative stress resistance, fungicide sensitivity, and virulence of Alternaria alternata. Appl. Environ. Microbiol. 2018, 84, e00086-18. [Google Scholar] [CrossRef] [Green Version]
  38. Ma, H.; Zhang, B.; Gai, Y.; Sun, X.; Chung, K.R.; Li, H. Cell-wall-degrading enzymes required for virulence in the host selective toxin-producing necrotroph Alternaria alternata of Citrus. Front. Microbiol. 2019, 10, 2514. [Google Scholar] [CrossRef]
  39. Fu, H.; Chung, K.R.; Gai, Y.; Mao, L.; Li, H. The basal transcription factor II H subunit Tfb5 is required for stress response and pathogenicity in the tangerine pathotype of Alternaria alternata. Mol. Plant Pathol. 2020, 21, 1337–1352. [Google Scholar] [CrossRef]
  40. Huang, F.; Fu, Y.; Nie, D.; Stewart, J.E.; Peever, T.L.; Li, H. Identification of a novel phylogenetic lineage of Alternaria alternata causing citrus brown spot in China. Fungal Biol. 2015, 119, 320–330. [Google Scholar] [CrossRef]
  41. Chung, K.-R.; Shilts, T.; Li, W.; Timmer, L.W. Engineering a genetic transformation system for Colletotrichum acutatum, the causal fungus of lime anthracnose and postbloom fruit drop of citrus. FEMS Microbiol. Lett. 2002, 213, 33–39. [Google Scholar] [CrossRef] [PubMed]
  42. Kohmoto, K.; Scheffer, R.P.; Whiteside, J.O. Host-selective toxins from Alternaria citri. Phytopathology 1979, 69, 667–671. [Google Scholar] [CrossRef] [Green Version]
  43. Masunaka, A.; Tanaka, A.; Tsuge, T.; Peever, T.L.; Timmer, L.W.; Yamamoto, M.; Yamamoto, H.; Akimitsu, K. Distribution and characterization of AKT homologs in the tangerine pathotype of Alternaria alternata. Phytopathology 2000, 90, 762–768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Nakashima, T.; Ueno, T.; Fukami, H.; Taga, T.; Masuda, H.; Osaki, K.; Otani, H.; Kohmoto, K.; Nishimura, S. Isolation and structures of AK-Toxin I and II, host-specific phytotoxic metabolites produced by Alternaria alternata Japanese pear pathotype. Agric. Biol. Chem. 1985, 49, 807–815. [Google Scholar] [CrossRef] [Green Version]
  45. Kohmoto, K.; Otani, H.; Kodama, M.; Nishimura, S. Host recognition: Can accessibility to fungal invasion be induced by host-specific toxins without necessitating necrotic cell death? Phytotoxins Plant Pathog. 1989, 27, 249–265. [Google Scholar]
  46. Gardner, J.; Mansour, I.S.; Scheffer, R. Effects of the host-specific toxin of Periconia circinata on some properties of sorghum plasma membranes. Physiol. Plant Pathol. 1972, 2, 197–206. [Google Scholar] [CrossRef]
  47. Wang, H.; Li, H.; Zhu, Z.; Wang, X. Expression of Ht2-related genes in response to the HT-Toxin of Exserohilum turcicum in Maize. Ann. Appl. Biol. 2010, 156, 111–120. [Google Scholar] [CrossRef]
  48. Okuno, T.; Ishita, Y.; Sawai, K.; Matsumoto, T. Characterization of alternariolide, a host-specific toxin produced by Alternaria mali roberts. Chem. Lett. 1974, 3, 635–638. [Google Scholar] [CrossRef]
  49. Akimitsu, K.; Peever, T.L.; Timmer, L.W. Molecular, ecological and evolutionary approaches to understanding Alternaria diseases of citrus. Mol. Plant Pathol. 2003, 4, 435–446. [Google Scholar] [CrossRef]
  50. Schröter, H.; Novacky, A.; Macko, V. Effect of Helminthosporium sacchari-toxin on cell membrane potential of susceptible sugarcane. Physiol. Plant Pathol. 1985, 26, 165–174. [Google Scholar] [CrossRef]
  51. Kohmoto, K.; Akimitsu, K.; Otani, H. Correlation of resistance and susceptibility of citrus to Alternaria alternata with sensitivity to host-specific toxins. Phytopathology 1991, 81, 719–722. [Google Scholar] [CrossRef]
  52. Maekawa, N.; Yamamoto, M.; Nishimura, S.; Kohmoto, K.; Kuwada, M.; Watanabe, Y. Studies on host-specific AF-toxins produced by Alternaria alternata strawberry pathotype causing Alternaria black spot of strawberry (1) production of host-specific toxins and their biological activities. Jpn. J. Phytopathol. 1984, 50, 600–609. [Google Scholar] [CrossRef]
  53. Slavov, S.; Mayama, S.; Atanassov, A. Toxin production of Alternaria alternata tobacco pathotype. Biotechnol. Biotechnol. Equip. 2004, 18, 90–95. [Google Scholar] [CrossRef]
  54. Walton, J.D.; Earle, E.D.; Gibson, B.W. Purification and structure of the host-specific toxin from Helminthosporium carbonum race 1. Biochem. Biophys. Res. Commun. 1982, 107, 785–794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Ciuffetti, L.M.; Manning, V.A.; Pandelova, I.; Betts, M.F.; Martinez, J.P. Host-selective toxins, Ptr ToxA and Ptr ToxB, as necrotrophic effectors in the Pyrenophora tritici-repentis–wheat interaction. New Phytol. 2010, 187, 911–919. [Google Scholar] [CrossRef]
  56. Liakopoulou-Kyriakides, M.; Lagopodi, A.; Thanassoulopoulos, C.; Stavropoulos, G.S.; Magafa, V. Isolation and synthesis of a host-selective toxin produced by Alternaria alternata. Phytochemistry 1997, 45, 37–40. [Google Scholar] [CrossRef]
  57. Bobylev, M.M.; Bobyleva, L.I.; Strobel, G.A. Synthesis and bioactivity of analogs of maculosin, a host-specific phytotoxin produced by Alternaria alternata on spotted knapweed (Centaurea maculosa). J. Agric. Food Chem. 1996, 44, 3960–3964. [Google Scholar] [CrossRef]
  58. Bottini, A.T.; Gilchrist, D.G. Phytotoxins. I. A 1-aminodimethylheptadecapentol from Alternaria alternata f. sp. lycopersici. Tetrahedron Lett. 1981, 22, 2719–2722. [Google Scholar] [CrossRef]
  59. Johnson, R.; Johnson, L.; Itoh, Y.; Kodama, M.; Otani, H.; Kohmoto, K. Cloning and characterization of a cyclic peptide synthetase gene from Alternaria alternata apple pathotype whose product is involved in AM-toxin synthesis and pathogenicity. Mol. Plant-Microbe Interact. 2000, 13, 742–753. [Google Scholar] [CrossRef] [Green Version]
  60. Yakimova, E.T.; Yordanova, Z.P.; Slavov, S.; Kapchina-Toteva, V.M.; Woltering, E.J. Alternaria alternata AT toxin induces programmed cell death in tobacco. J. Phytopathol. 2009, 157, 592–601. [Google Scholar] [CrossRef]
  61. Stierle, A.C.; Cardellina, J.H.; Strobel, G.A. Maculosin, a host-specific phytotoxin for spotted knapweed from Alternaria alternata. Proc. Natl. Acad. Sci. USA 1988, 85, 8008–8011. [Google Scholar] [CrossRef] [Green Version]
  62. Lin, C.H.; Yang, S.L.; Chung, K.R. The YAP1 homolog–mediated oxidative stress tolerance is crucial for pathogenicity of the necrotrophic fungus Alternaria alternata in citrus. Mol. Plant-Microbe Interact. 2009, 22, 942–952. [Google Scholar] [CrossRef] [PubMed]
  63. Yang, S.L.; Yu, P.L.; Chung, K.R. The glutathione peroxidase-mediated reactive oxygen species resistance, fungicide sensitivity and cell wall construction in the citrus fungal pathogen Alternaria alternata. Environ. Microbiol. 2016, 18, 923–935. [Google Scholar] [CrossRef]
  64. Gai, Y.; Liu, B.; Ma, H.; Li, L.; Chen, X.; Moenga, S.; Riely, B.; Fayyaz, A.; Wang, M.; Li, H. The methionine biosynthesis regulator AaMetR contributes to oxidative stress tolerance and virulence in Alternaria alternata. Microbiol. Res. 2019, 219, 94–109. [Google Scholar] [CrossRef]
  65. Yu, P.L.; Chen, L.H.; Chung, K.R. How the pathogenic fungus Alternaria alternata copes with stress via the response regulators SSK1 and SHO1. PLoS ONE 2016, 11, e0149153. [Google Scholar] [CrossRef] [Green Version]
  66. Wu, P.C.; Choo, C.Y.L.; Lu, H.Y.; Wei, X.Y.; Chen, Y.K.; Yago, J.I.; Chung, K.R. Pexophagy is critical for fungal development, stress response, and virulence in Alternaria alternata. Mol. Plant Pathol. 2022, 23, 1538–1554. [Google Scholar] [CrossRef]
  67. Chung, K.R.; Wu, P.C.; Chen, Y.K.; Yago, J.I. The siderophore repressor SreA maintains growth, hydrogen peroxide resistance, and cell wall integrity in the phytopathogenic fungus Alternaria alternata. Fungal Genet. Biol. 2020, 139, 103384. [Google Scholar] [CrossRef]
  68. Chung, K.-R. Stress response and pathogenicity of the necrotrophic fungal pathogen Alternaria alternata. Scientifica 2012, 2012, 635431. [Google Scholar] [CrossRef] [Green Version]
  69. Meena, M.; Samal, S. Alternaria host-specific (HSTs) toxins: An overview of chemical characterization, target sites, regulation and their toxic effects. Toxicol. Rep. 2019, 6, 745–758. [Google Scholar] [CrossRef]
  70. Lin, C.H.; Yang, S.L.; Chung, K.R. Cellular responses required for oxidative stress tolerance, colonization, and lesion formation by the necrotrophic fungus Alternaria alternata in citrus. Curr. Microbiol. 2011, 62, 807–815. [Google Scholar] [CrossRef]
  71. Wang, H.; Guo, Y.; Luo, Z.; Gao, L.; Li, R.; Zhang, Y.; Kalaji, H.M.; Qiang, S.; Chen, S. Recent advances in Alternaria phytotoxins: A review of their occurrence, structure, bioactivity, and biosynthesis. J. Fungi 2022, 8, 168. [Google Scholar] [CrossRef] [PubMed]
  72. Gardner, J.; Kono, Y.; Chandler, J. Bioassay and host-selectivity of Alternaria citri toxins affecting rough lemon and mandarins. Physiol. Mol. Plant Pathol. 1986, 29, 293–304. [Google Scholar] [CrossRef]
  73. Solel, Z.; Kimchi, M. Susceptibility and resistance of citrus genotypes to Alternaria alternata pv. citri. J. Phytopathol. 1997, 145, 389–391. [Google Scholar] [CrossRef]
  74. Peever, T.L.; Olsen, L.; Ibañez, A.; Timmer, L.W. Genetic differentiation and host specificity among populations of Alternaria spp. causing brown spot of grapefruit and tangerine × grapefruit hybrids in florida. Phytopathology 2000, 90, 407–414. [Google Scholar] [CrossRef] [Green Version]
  75. Vicent, A.; Badal, J.; Asensi, M.J.; Sanz, N.; Armengol, J.; García-Jiménez, J. Laboratory evaluation of citrus cultivars susceptibility and influence of fruit size on fortune mandarin to infection by Alternaria alternata pv. citri. Eur. J. Plant Pathol. 2004, 110, 245–251. [Google Scholar] [CrossRef]
Figure 1. Different citrus accessions display distinct sensitivity to ACT-toxin. (A) Isolation procedure to obtain ACT crude extract. (B) Virulence test of ACT crude extract on leaves of different citrus accessions.
Figure 1. Different citrus accessions display distinct sensitivity to ACT-toxin. (A) Isolation procedure to obtain ACT crude extract. (B) Virulence test of ACT crude extract on leaves of different citrus accessions.
Agronomy 12 03181 g001
Figure 2. ACTT5 knockout and virulence assay. (A) Schematic diagram for ACTT5 knockout. (B) ACTT5 was not involved in vegetative growth. (C) ACTT5 was required for virulence. “S” means “Sugar belle”, “M” means “Minneola”, “C” means “Clementina”, “D” means “Dancy”.
Figure 2. ACTT5 knockout and virulence assay. (A) Schematic diagram for ACTT5 knockout. (B) ACTT5 was not involved in vegetative growth. (C) ACTT5 was required for virulence. “S” means “Sugar belle”, “M” means “Minneola”, “C” means “Clementina”, “D” means “Dancy”.
Agronomy 12 03181 g002
Figure 3. ACTT5 is not involved in conidiation, vegetative growth, and multi-stress resistance. (A) Microscopic observation of mycelium of wild-type (WT) and ∆∆ACTT5. (B) Conidiation of wild-type and ∆∆ACTT5. (C) Multi-stress analysis of wild-type and ∆∆ACTT5. (D) The percentage of growth reduction determined by comparing a cumulative percentage of the growth of Z7 and ∆∆ACTT5 grown on same medium is also shown.
Figure 3. ACTT5 is not involved in conidiation, vegetative growth, and multi-stress resistance. (A) Microscopic observation of mycelium of wild-type (WT) and ∆∆ACTT5. (B) Conidiation of wild-type and ∆∆ACTT5. (C) Multi-stress analysis of wild-type and ∆∆ACTT5. (D) The percentage of growth reduction determined by comparing a cumulative percentage of the growth of Z7 and ∆∆ACTT5 grown on same medium is also shown.
Agronomy 12 03181 g003
Table 1. Examples of HSTs.
Table 1. Examples of HSTs.
NoPathogenToxinHostReference
1A. alternata Japanese pear pathotypeAKPear[44,45]
2Helminthosporium victoriaeHVOat[4]
3Periconia circinataPCSorghum[46]
4Helminthosporium turcicumHTMaize[47]
5A. alternata apple pathotypeAMApple[48,49]
6A. alternata tangerine pathotypeACTCitrus[2,49]
7Helminthosporium maydisHMTMaize[5]
8Helminthosporium sacchariHSSugarcane[50]
9A. alternata tomato pathotypeAALTomato[18]
10A. alternata rough lemon pathotypeACRRough lemon[42,51]
11A. alternata strawberry pathotypeAFStrawberry[52]
12A. alternata tobacco pathotypeATTobacco[53]
13Helminthosporium carbonumHCMaize[54]
14Pyrenophora tritici-repentisPTRWheat[55]
15A. alternata sunflower pathotypeAS-ISunflower[56]
16A. alternata spotted knapweed pathotypeMaculosinKnapweed[57]
17A. brassicaeABRBrassica spp.[3]
Table 2. HSTs produced by A. alternata.
Table 2. HSTs produced by A. alternata.
NoPathotypeHSTsDiseaseChemical CharacteristicsReference
1Tomato pathotypeAALAlternaria stem canker of tomatoAminopentol esters[1,18,58]
2Tangerine pathotypeACTCitrus brown spotEpoxy-decatrienoic esters[2]
3Rough lemon pathotypeACRLeaf spot of rough lemonTerpenoid[19]
4Strawberry pathotypeAFBlack spot of strawberryEpoxy-decatrienoic esters[17]
5Japanese pear pathotypeAKBlack spot of Japanese pearEpoxy-decatrienoic esters[16]
6Apple pathotypeAMAlternaria blotch of appleCyclic peptide[59]
7Sunflower pathotypeAS-ILeaf spot of sunflowerTetrapeptide[56]
8Tobacco pathotypeATBrown spot of tobacco---[60]
9Spotted knapweed pathotypeMaculosinBlack leaf blight of knapweedTetrapeptide[61]
Table 3. Key genes essential for ACT biosynthesis.
Table 3. Key genes essential for ACT biosynthesis.
NoGeneNameCopy NumberAccession NumberReference
1ACTT2Hydrolase≥2AALT_g11743[20]
2ACTTS2Enoyl-reductase≥2AALT_g12031[22]
3ACTT3HMG-CoA hydrolase≥2AALT_g11755[25]
4ACTTS3Polyketide synthase≥3AALT_g11750[27]
5ACTT5Acyl-CoA synthetase≥3AALT_g11751[24]
6ACTT6Enoyl-CoA hydratase2AALT_g12047[24]
7ACTTRZn(II)2Cys6 transcription factor≥2AALT_g11754[28]
Table 4. ROS detoxification-related genes in A. alternata tangerine pathotype.
Table 4. ROS detoxification-related genes in A. alternata tangerine pathotype.
NoGeneVegetative GrowthConidiationPathogenicityAccession NumberReference
1Hog1RequiredRequiredRequiredGQ414509[35]
2Skn7RequiredRequiredRequiredJQ716919[36]
3Ap1Required---RequiredFJ376607[62]
4Gpx3RequiredRequiredRequiredACY73852[63]
5Tsa1Not requiredNot requiredRequiredMG593564[37]
6Trr1RequiredRequiredRequiredMG593563[37]
7Glr1RequiredRequiredRequiredMG593559[37]
8NoxARequiredRequiredRequiredJN900389[32]
9NoxBRequiredRequiredRequiredJX136700[32]
10NoxRRequiredRequiredRequiredJX207117[32]
11MetRRequiredRequiredRequiredAa03030[64]
12SSK1RequiredRequiredRequiredKU170060[65]
13Tbf1RequiredRequiredRequiredMT184174[39]
14Atg8RequiredRequiredRequiredOK617334[66]
15SreARequiredRequiredRequiredOWY49902.1[67]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Huang, S.; Jia, Z.; Li, H.; Zhang, S.; Shen, J.; Gai, Y.; Jiao, C.; Sun, X.; Duan, S.; Wang, M.; et al. ACT-Toxin, the Key Effector for the Virulence of Alternaria alternata Tangerine Pathotype to Specific Citrus Species. Agronomy 2022, 12, 3181. https://doi.org/10.3390/agronomy12123181

AMA Style

Huang S, Jia Z, Li H, Zhang S, Shen J, Gai Y, Jiao C, Sun X, Duan S, Wang M, et al. ACT-Toxin, the Key Effector for the Virulence of Alternaria alternata Tangerine Pathotype to Specific Citrus Species. Agronomy. 2022; 12(12):3181. https://doi.org/10.3390/agronomy12123181

Chicago/Turabian Style

Huang, Suya, Zhaohui Jia, Hangfei Li, Shuting Zhang, Junying Shen, Yunpeng Gai, Chen Jiao, Xuepeng Sun, Shuo Duan, Min Wang, and et al. 2022. "ACT-Toxin, the Key Effector for the Virulence of Alternaria alternata Tangerine Pathotype to Specific Citrus Species" Agronomy 12, no. 12: 3181. https://doi.org/10.3390/agronomy12123181

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop